• Keine Ergebnisse gefunden

Due to the fact that most deep-sea benthic species are deposit feeders (Sanders and Hessler, 1969), the locally qualitatively and quantitatively, variable import of POM from the ocean’s surface waters plays a crucial role for macro-, meio-, and microorganisms living in deep surface sediments (Gooday and Turley, 1990). Mainly consisting of phytoplankton, marine snow, fecal pellets, (dead) zooplankton and molts, this material undergoes different steps of degradation during its passage from the photic, epipelagic (0–200 mbsl), through the mesopelagic (200–1,000 mbsl) to the actual deep-sea zones, in particular the bathypelagic (1,000–4,000 mbsl), the abyssal (4,000–6,000 mbsl), and the hadal zone (6,000–11,000 mbsl).

Depending on the residence time in the water column, the bioavailable part of POM finally reaching the deep-seafloor may be small (De La Rocha and Passow, 2007 and references therein). The refractory remainders such as animal skeletons are continuously accumulating at the seafloor and turn into deeply buried sediment over time, thereby representing the largest global reservoir organic carbon (Parkes et al., 2000 and references therein).

2.4DEEP-SEA SEDIMENT TYPES

Grain size (Gray, 1974) and sediment heterogeneity (Etter and Grassle, 1992) may additionally govern community composition and distribution of macro-, meio-, and microorganisms in deep-sea benthic environments. In relation to their basic sources, deep-sea sediments may be biogenic (POM from pelagic primary production, benthic in-situ production), lithogenous (terrestrial weathering and transport by wind and rivers), hydrogenous (precipitation from seawater or pore water), volcanic, or cosmic (Seibold and Berger, 1996). According to grain size and settling velocity, lithogenous gravel and sandy fractions usually are deposited along the coast, while silt and clay are transported farther offshore through waves and currents, hence dominating the basically biogenous deep-sea

sediments. Regional deviations may be linked to currents, downslope slides, submarine canyon dynamics, or to a release of ice-trapped rock material in polar waters (e.g., Ramseier et al., 2001). Covering almost one-half of the shelves and more than half of the deep ocean bottom, biogenous sediments mainly consist of calcite, aragonite, opal, and calcium phosphate, originating from foraminifera, diatoms, and radiolarians (Hay et al., 1988).

2.5DEEP BIOSPHERE OF DEEP-SEA SEDIMENTS

Microbiological studies on sediment cores collected during several cruises of the Deep Sea Drilling Project (DSDP), the Ocean Drilling Program (ODP), and the Integrated Ocean Drilling Program (IODP) gave evidence for the presence of complex microbial communities in deeply buried marine sediments down to several hundred meters below seafloor (e.g., Whelan et al., 1986; Parkes et al., 1994; Roussel et al., 2008). Most striking, new insights into subsurface microbiology were gained during the ODP cruise Leg 201 to the equatorial Pacific Ocean and the continental margin of Peru, including sites recognized as most typical for oceanic subsurface environments (D’Hondt et al., 2002). A large fraction of the sub-seafloor bacteria has been proven to be alive and culturable, displaying turnover rates (based on sulfate reduction as dominating mineral process at these sites) comparable to surface sediment communities (D’Hondt et al., 2004; Schippers et al., 2005).

After a logarithmic decline within the uppermost 6 meters below seafloor (mbsf) (Parkes et al., 1994) to about 40 mbsf (Schippers et al., 2005), bacterial cells have proven to be more or less evenly distributed down to several hundred mbsf. Local peaks within these deeply buried sediments seem to mirror sulfate (diffusing from crustal fluids) and methane (from in-situ production) concentration shifts (Engelen et al., 2008). However, published variations of absolute cell numbers (by a factor of up to 3) have to be treated with caution: varying estimations not only depend on the geochemical conditions at the respective sampling sites,

but also on the enumeration techniques applied. Calculations based on early results revealed that sub-seafloor sediments comprise –at least – half of all prokaryotic cells and up to one-third of the living biomass on Earth pointing to a slow-growing strategy of high biomass in areas of low-energy flux (Whitman et al., 1998).

The prokaryotic community in deeply buried sediments can not exclusively be traced back to contaminations from biologically active surface layers or reactivation of spores and dormant cells (Parkes et al., 2000 and references therein).

Porewater chemistry data obtained from sites throughout the world’s oceans (ODP, DSDP) showed that sulfate reduction, methanogenesis, and fermentation are the principal degradative metabolic processes in subsurface sediments. These results give evidence for significant lower metabolic rates for the subsurface compared to the surface biosphere and for methanogenesis becoming more important the more sulfate gets depleted with increasing sediment depth (D’Hondt et al., 2002 and references therein).

2.6WINDOWS TO THE SUBSURFACE BIOSPHERE

2.6.1 Hydrothermal vents

The discovery of the “ocean vents” near Galapagos Island (Corliss et al., 1979) was the first proof for the active movement of the gigantic oceanic plates of the Earth’s crust creating series of cracks in the ocean floor, teeming with life. At these discharge areas, hydrothermal fluids with temperatures of more than 400°C (Haase et al., 2007) mix up with the cold ocean seawater, resulting in a precipitation of dissolved metals and in the formation of characteristic chimneys over time. Iron and sulfide precipitates turn the smokers black (“black smokers,”

Figure 1), while barium, calcium, and silicon minerals result in “white smokers.” Thermal precipitation and/or direct magma degassing of H2, H2S, CH4, CO, and CO2 in combination with oxygen as electron acceptor provide enough energy to support a highly productive and

physiologically diverse chemoautotrophic microbial community (Reysenbach and Shock, 2002).

As the highly diverse and dense hot vent macrofauna (e.g., vestimentiferan tubeworms, bivalve mollusks, provannid gastropods, alvinellid polychaete, and bresiliid shrimps) cannot feed on the released chemicals themselves, they either feed on chemoautotrophic microbes or host them as symbionts. The predominant endosymbionts are mesophilic to moderately thermophilic chemoautotrophs (mostly Gammaproteobacteria), whereas most episymbionts belong to the Epsilonproteobacteria, which can oxidize H2 and sulfur compounds while reducing oxygen, nitrate, and sulfur compounds (for review, see Nakagawa and Takai, 2008).

The vent habitat proved to harbor methanogens (Methanococcus), sulfate-reducers (Archaeoglobus), and facultative autotrophs and heterotrophs such as the thermophilic aerobic Thermus and Bacillus (Harmsen et al., 1997) or, for example, Thermococcus, Phyrococcus, Desulfurococcus (Prieur et al., 1995; Teske et al., 2000; Nercessian et al., 2003; Schrenk et al., 2003). Generally detected archaeal phylotypes were affiliated with hyperthermophilicCrenarchaeota, Euryarchaeota Group I, II, III (Takai and Horikoshi, 1999) and the “Deep-sea Hydrothermal Vent Euryarchaeotal Group” (Hoek et al., 2003).

2.6.2 Cold seeps and mud volcanoes

Just a few years after the discovery of the hydrothermal vent systems, cold seep ecosystems were reported from active and passive continental margins and subduction zones all over the world (Aharon, 1994 and references therein).

High-pressure, low oxygen and low-temperature conditions favour the formation of marine gas hydrates. In the subsurface realm, such gas reservoirs are stored in a crystalline form, whereas they get, dissolved in pore waters and finally leave the sediment surface in gaseous form. High fluxes of methane, sulfide, and other reduced elements characterize these

ecosystems such as cold seeps, hydrocarbon vents and mud volcanoes, often leaving mineral precipitation in their immediate surroundings. Coupled to sulfate reduction, rich bacterial and archaeal communities perform anaerobic oxidation of hydrocarbons, but predominately of methane (Boetius et al., 2000; Borowski et al., 2000; Treude et al., 2005). Conversion of methane is mainly mediated by two different groups of anaerobic methanotrophic archaea (ANME-I and ANME-II) (Nauhaus et al., 2005), forming syntrophic consortia with the sulfate-reducing bacteria (SRB) Desulfosarcina and Desulfococcus (Hinrichs et al., 1999;

Boetius et al., 2000; Michaelis et al., 2002; Knittel et al., 2003).

The methane-emitting Haakon Mosby Mud Volcano (HMMV, Barents Sea) has shown to harbor three key communities in methane conversion such as aerobic, methanotrophic bacteria (Methylococcales), anaerobic methanotrophic archaea (ANME-2) thriving below siboglinid tubeworms, and a previously undescribed clade of archaea (ANME-3) associated with bacterial mats (Niemann et al., 2006). Similarly, some cold seeps on the deeper Black Sea shelf are characterized by intense methane bubble discharge, mainly related to microbial methanogenesis (Pape et al., 2008 and references therein). Diffuse gas seeps in more shallow, oxic Black Sea waters often exhibit a netlike coverage of microbial mats similar to Beggiatoa-mats observed at HMMV (Figure 2). Beggiatoa spp. are discussed as keystone members of seep communities owing to their ability to (directly and indirectly) influence the metabolic activity of d-Proteobacteria, Planctomycetales, and ANME archaea by providing sulfate and ammonia as reactants (Mills et al., 2004).

The question remains, to which extent such seep systems influence the global methane cycle, as the quantification of bubble dissolution and/or the release of methane- rich pore fluids from the sediment into the hydrosphere is difficult to achieve (Vogt et al., 1999; Reeburgh, 2007).

Niemann et al. (2006) estimated that methanotrophy at active marine mud volcanoes

consumes less than 40% of the total methane flux, due to limitations of the relevant electron acceptors in the upward flowing, sulphate- and oxygen-free fluids.

2.7DEEP BIOSPHERE OF THE OCEANIC CRUST

The fact that microorganisms are present in the subsurface realm had been reported decades ago in terrestrial subsurface environments (Farrell and Turner, 1931; Lipman, 1931). Early drilling operations performed for commercial purposes such as mining, oil and hot water recovery, and the search for underground waste repositories reported on the existence of a large community of microorganisms obviously involved in geochemical processes in the deep biosphere (Gold, 1992; Pedersen, 1993 and references therein). Hence, it was only in the early 1990s that scientists started to focus on the investigation of prospering life beneath the Earth’s crust, thanks to a chance encounter of a deep oceanic, volcanic eruption during a dive onboard the submersible Alvin, releasing white microbial bulk mats (Haymon et al., 1993). The upper layers of the oceanic crust are characterized by high basaltic porosity, hosting a vast hydrothermal reservoir (Johnson and Pruis, 2003) inhabited by a microbial community composed of species that are also found in deep-sea waters, sediments, and the deep oceanic crust (Thorseth et al., 2001; Huber et al., 2006). Among the most prominent anaerobic thermophiles indigenous for the oceanic crust, the Ammonifex group of bacteria (Nakagawa et al., 2006) or groups within Crenarchaeota, Euryarchaeota, and Korarchaeota (Ehrhardt et al., 2007). Since 3.5 billion years, basaltassociated glass textures and vesicular cavities within the basaltic matrix provide niches for microbial colonization (Furnes et al., 2004; Peckmann et al., 2008). For instance, the fossil record of the oceanic crust even gives evidence for a previous fungal life in deep ocean basaltic rocks (Schumann et al., 2004).

Much effort has been put into the investigation of the deep biosphere of the deep sea during the past 20 years. However, we still are neither aware of the final composition of the living

subsurface community, nor of its interrelationship to, for example, crustal fluid-derived compounds, nor of its global impact.

2.8FIGURES

Figure 1

Black smoker “Candelabra,” Logachev hydrothermal field, Mid-Atlantic-Ridge (3,000 m water depth). (By courtesy of MARUM, Center for Marine Environmental Sciences, Bremen, Germany.)

Figure 2

Shallowwater seep area “GHOSTDABS-field,” Ukrainian shelf, Black Sea. White nets are constructed of sulfide oxidizing bacteria (“Beggiatoa”). (By courtesy of Karin Hissmann and Jürgen Schauer, Jago Team, Leibniz-Institut für Meereswissenschaften (IFM-GEOMAR) Kiel, Germany.)

2.9 REFERENCES

Aharon P. 1994. Geology and biology of modern and ancient submarine hydrocarbon seeps and vents: An introduction. Geo-Marine Letters, 14(2): 69-73.

Boetius A, Ravenschlag K, Schubert CJ, Rickert D, Widdel F, Gieseke A, Amann R, Jorgensen B B, Witte U, Pfannkuche O. 2000. A marine microbial consortium apparently mediating anaerobic oxidation of methane. Nature, 407(6804): 623-626.

Borowski WS, Cagatay N, Ternois Y, Paull CK. 2000. Data report: carbon isotopic composition of dissolved CO2, CO2 gas, and methane, Blake-Bahama Ridge and northeast Bermuda Rise, ODP Leg 172. Proceedings of the Ocean Drilling Program, Scientific Results 172. http://www.odp.tamu.edu/publications/172_SR/chap_03/chap_

03.htm

Corliss JB, Dymond J, Gordon LI, Edmond JM, von Herzen RP, Ballard RD, Green K, Williams D, Bainbridge A, Crane K, VanAndel TH. 1979. Submarine Thermal Springs on the Galápagos Rift. Science, 203: 1073-1083.

Danovaro R, Della Croce N, Dell'Anno A, Mauro F, Marrale D, Martorano D. 2000. Seasonal changes and biochemical composition of the labile organic matter flux in the Cretan Sea.

Progress In Oceanography, 46: 259-278.

De La Rocha C, Passow U. 2007. Factors influencing the sinking of POC and the efficiency of the biological carbon pump. Deep Sea Research II, 54: 639-658.

D'Hondt S, Rutherford S, Spivack AJ. 2002. Metabolic Activity of Subsurface Life in Deep-Sea Sediments. Science, 295(5562): 2067-2070.

D'Hondt S, Jorgensen BB, Miller DJ, Batzke A, Blake R, Cragg BA, Cypionka H, Dickens GR, Ferdelman T, Hinrichs KU, Holm NG, Mitterer R, Spivack A, Wang G, Bekins B, Engelen B, Ford K, Gettemy G, Rutherford SD, Sass H, Skilbeck CG, Aiello IW, Guerin G, House CH, Inagaki F, Meister P, Naehr T, Niitsuma S, Parkes RJ, Schippers A, Smith DC, Teske A, Wiegel J, Padilla CN, Acosta JLS. 2004. Distributions of Microbial Activities in Deep Subseafloor Sediments. Science, 306(5705): 2216-2221.

Engelen B, Ziegelmüller K, Wolf L, Köpke B, Gittel A, Cypionka H, Treude T, Nakagawa S, Inagaki F, Lever MA, Steinsbu BO. 2008b. Fluids from the Oceanic Crust Support Microbial Activities within the Deep Biosphere. Geomicrobiology Journal, 25(1): 56 - 66.

Ehrhardt CJ, Haymon RM, Lamontagne MG, Holden PA. 2007. Evidence for hydrothermal Archaea within the basaltic flanks of the East Pacific Rise. Environmental Microbiology, 9: 900–912.

Etter RJ, Grassle JF. 1992. Patterns of species diversity in the deep sea as a function of sediment particle size diversity. Nature, 360: 576-578.

Farrell MA, Turner HG. 1931. Bacteria in anthracite coal. Journal of Bacteriology, 23: 155pp.

Furnes H, Banerjee NR, Muehlenbachs K, Staudigel H, de Wit M. 2004. Early Life Recorded in Archean Pillow Lavas. Science, 304: 578- 581.

Gage JD, Tyler PA. 1991. sea Biology - A Natural History of Organsims at the Deep-sea Floor. Cambridge: Cambridge University Press.

Gold T. 1992. The deep, hot biosphere. Proceedings of the National Academy of Sciences of the United States of America, 89 (13): 6045-6049.

Gooday AJ, Turley CM. 1990. Responses by Benthic Organisms to Inputs of Organic Material to the Ocean Floor: A Review. Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences, 331(1616): 119-137.

Gray JS. 1974. Animal-sediment relationships. Oceanography and Marine Biology: an Annual Review, 12: 223-261.

Haase KM, Petersen S, Koschinsky A, Seifert R, Devey CW, Keir R, Lackschewitz KS, Melchert B, Perner M, Schmale O, Süling J, Dubilier N, Zielinski F, Fretzdorff S, Garbe-Schönberg D, Westernströer U, German CR, Shank TM, Yoerger D, Giere O, Kuever J, Marbler H, Mawick J, Mertens C, Stöber U, Walter M. 2007. Young volcanism and related hydrothermal activity at 5°S on the slow-spreading southern Mid-Atlantic Ridge.

G3 8 (11):1-17.

Harmsen HJM., Prieur D, Jeanthoni C. 1997. Distribution of Microorganisms in Deep-Sea Hydrothermal Vent Chimneys Investigated by Whole-Cell Hybridization and Enrichment Culture of Thermophilic Subpopulations. Applied and Environmental Microbiology, 63 (7): 2876–2883.

Hay WW, Sloan II JL, Wold CN. 1988. Mass/age distribution and composition of sediments on the ocean floor and the global rate of sediment subduction. Journal of Geophysical research, 93 (B12): 14.933–14.940.

Haymon R, Fornari D, Von Damm K, Lilley M, Perfit M, Edmond J, Shanks WC, W III, R.

Lutz, J. Grebmeier, S. Carbotte, D. Wright, E. McLaughlin, M. Smith, N. Beedle, E.

Olson, 1993. Volcanic eruption of the mid-ocean ridge along the East Pacific Rise at 9º45-52'N: Direct submersible observation of seafloor phenomena associated with an eruption event in April, 1991. Earth and Planetary Science Letters, 119:85-101.

Hinrichs KU, Hayes JM, Sylva SP, Brewer PG, DeLong EF. 1999. Methane-consuming archaebacteria in marine sediments. Nature 398: 802–805.

Hoek J, Banta A, Hubler F, Reysenbach AL. 2003. Microbial diversity of a sulphide spire located in the Edmond deep-sea hydrothermal vent field on the Central Indian Ridge.

Geobiology,1: 119–127.

Huber JA, Johnson HP, Butterfield DA, Baross JA. 2006. Microbial life in ridge flank crustal fluids. Environmental Microbiology, 8: 88–99.

Johnson, HP, Pruis, MJ, 2003. Fluxes of fluid and heat from the oceanic crustal reservoir.

Earth and Planetary Science Letters, 216: 565–574.

Knittel K, Boetius A, Lemke A, Eilers H, Lochte K, Pfannkuche O, Linke K. 2003. Activity, Distribution, and Diversity of Sulfate Reducers and Other Bacteria in Sediments above Gas Hydrate (Cascadia Margin, Oregon). Geomicrobiology Journal, 20: 269–294.

Lauro F, Bartlett D. 2008. Prokaryotic lifestyles in deep sea habitats. Extremophiles 12:15–

25.

Levin LA, Etter RJ, Rex MA, Gooday AJ, Smith CR, Pineda J, Stuart CT, Hessler RR, Pawson DL. 2001. Environmental influences on regional deep-sea species diversity.

Annual Review of Ecology and Systematics, 32: 51–93.

Lipman CB. 1931. Living micro-organisms in ancient rocks. Journal of Bacteriology, 22:

183pp.

Lochte K, Turley CM. 1988. Bacteria and cyanobacteria associated with phytodetritus in the deep sea. Nature, 333: 67-68.

Michaelis W, Seifert R, Nauhaus K, Treude T, Thiel V, Blumenberg M, Knittel K, Gieseke A, Peterknecht K, Pape T, Boetius A, Amann R, Jorgensen BB, Widdel F, Peckmann J, Pimenov NV, Gulin MB. 2002. Microbial Reefs in the Black Sea Fueled by Anaerobic Oxidation of Methane. Science, 297(5583): 1013-1015.

Nakagawa S, Inagaki F, Suzuki Y, Steinsbu BO, Lever MA, Takai K, Engelen B, Sako Y, Wheat CG, Horikoshi K. 2006. Microbial community in black rust exposed to hot ridge flank crustal fluids. Applied and Environmental Microbiology, 72(10): 6789-6799.

Nakagawa S, Takai K. 2008. Deep-sea vent chemoautotrophs: diversity, biochemistry and ecological significance. FEMS Microbiology Ecology, 65(1): 1-14.

Nauhaus K, Treude T, Boetius A, Krüger M. 2005. Environmental regulation of the anaerobic oxidation of methane: a comparison of ANME-I and ANME-II communities.

Environmental Microbiology, 7 (1): 98–106.

Nercessian O, Reysenbach AL, Prieur D, Jeanthon C. 2003. Archaeal diversity associated with in situ samplers deployed on hydrothermal vents on the East Pacific Rise (13 ∞N).

Environmental Microbiology, 5(6): 492–502.

Parkes RJ, Cragg BA, et al., 1994. Deep bacterial biosphere in Pacific Ocean sediments.

Nature, 371(6496): 410-413.

Parkes, J.R. et al. , 2000. Recent studies on bacterial populations and processes in subseafloor sediments: a review. Hydrogeology Journal, 8: 11–28.

Peckmann J, Bach W, Behrens K, Reitner J. 2008. Putative cryptoendolithic life in Devonian pillow basalt, Rheinisches Schiefergebirge, Germany. Geobiology, 6: 125–135.

Pedersen K. 1993. The deep subterranean biosphere. Earth-Science Reviews, 34(4): 243-260.

Pfannkuche O. 1992. Organic carbon flux through the benthic community in the temperate abyssal northeast Atlantic. In: Rowe, G.T., Pariente, V. (eds) Deep-sea food chains and the global carbon cycle. Kluwer Academic Publishers: The Netherlands, pp 183-198.

Prieur D, Erauso G, Jeanthon C. 1995. Hyperthermophilic life at deep-sea hydrothermal vents.

Planetary and Space Science, 43(1-2): 115-122.

Ramseier RO, Garrity C, Martin T. 2001. An overview of sea-ice conditions in the Greenland Sea and the relationship of oceanic sedimentation to the ice regime. In: Schäfer P., Ritzrau W., Schlüter M., Thiede J. (eds). The Northern North Atlantic – A changing Environment.

Berlin: Springer Verlag, pp 19-38.

Reeburgh WS. 2007. Oceanic Methane Biogeochemistry. Chemical Reviews, 107(2): 486-513.

Reysenbach AL, Shock E. 2002. Merging Genomes with Geochemistry in Hydrothermal Ecosystems. Science, 296(5570): 1077-1082.

Roussel EG, Cambon Bonavita MA, Querellou J, Cragg BA, Webster G, Prieur D, Parkes RJ.

2008. Extending the Sub–Sea-Floor Biosphere. Science, 320(5879): 1046.

Sanders HL, Hessler RR. 1969. Ecology of the Deep-Sea Benthos. Science, 163(3874): 1419-1424.

Schippers A, Neretin LN et al., 2005. Prokaryotic cells of the deep sub-seafloor biosphere identified as living bacteria. Nature, 433(7028): 861-864.

Schrenk MO, Kelley DS, Delaney JR, Baross JA. 2003. Incidence and diversity of microorganisms within the walls of an active deep-sea sulfide chimney. Applied and Environmental Microbiology, 69: 3580–3592.

Schumann G, Manz W, Reitner J,Lustrino M. 2004. Ancient Fungal Life in North Pacific Eocene Oceanic Crust. Geomicrobiology Journal, 21:241–246.

Seibold E, Berger WH. 1996. The sea floor - An introduction to marine geology. (3rd ed)

Teske A, Brinkhoff T, Muyzer G, Moser DP, Rethmeier J, Jannasch HW 2000. Diversity of Thiosulfate-Oxidizing Bacteria from Marine Sediments and Hydrothermal Vents. Applied and Environmental Microbiology, 66 (8): 3125–3133.

Thorseth IH, Torsvik T, Torsvik V, Daae FL, Pedersen RB. 2001. Diversity of life in ocean floor basalt. Earth and Planetary Science Letters, 194: 31–37.

Treude T, Knittel K, Blumenberg M, Seifert R, Boetius A. 2005. Subsurface Microbial Methanotrophic Mats in the Black Sea. Appl. Environ. Microbiol., 71(10): 6375-6378.

Turley CM, Mackie PJ. 1995. Bacterial and cyanobacterial flux to the deep NE Atlantic on sedimenting particles. Deep-Sea Research I, 42: 1453-1474.

Tyler PA. 2003. Ecosystems of the deep oceans. Ecosystems of the world, 28. Elsevier:

Amsterdam.

Vanucci S, Dell'Anno A, Pusceddu A, Fabiano M, Lampitt RS, Danovaro R. 2001. Microbial assemblages associated with sinking particles in the Porcupine Abyssal Plain (NE Atlantic Ocean). Progress in Oceanography, 50: 105-121.

Venter A. 2001. Deep cold biosphere. Trends in Microbiology, 9(7): 313-313.

Vorobyova E, Soina V. et al. ,1997. The deep cold biosphere: facts and hypothesis. FEMS Microbiology Reviews 20: 277-290.

Whelan JK, Oremland R, Tarafa M, Smith R, Howarth R, Lee C. 1986. Evidence for sulfate-reducing and methane-producing micro-organisms in sediments from Site 618, Site 619, and Site 622. Initial Reports of the Deep Sea Drilling Project, 96:767–775.

Whitman WB, Coleman DC, Wiebe WJ. 1998. Prokaryotes: The unseen majority. Proc. Natl.

Acad. Sci. USA, 95, 6578–6583.

C

HAPTER

3

MICROORGANISMS ISOLATED FROM DEEP SEA